Role of Endosomal Escape for Disulfide-Based Drug Delivery from

Aug 2, 2010 - γ-PGA-Coated Mesoporous Silica Nanoparticles with Covalently Attached Prodrugs for Enhanced Cellular Uptake and Intracellular GSH-Respo...
0 downloads 8 Views 3MB Size
pubs.acs.org/NanoLett

Role of Endosomal Escape for Disulfide-Based Drug Delivery from Colloidal Mesoporous Silica Evaluated by Live-Cell Imaging Anna M. Sauer,† Axel Schlossbauer,† Nadia Ruthardt, Valentina Cauda, Thomas Bein,* and Christoph Bra¨uchle* Department of Chemistry and Center for Nanoscience (CeNS), University of Munich (LMU), Butenandtstrasse 11, Gerhardt-Ertl-Building, 81377 Munich, Germany ABSTRACT Redox-driven intracellular disulfide-cleavage is a promising strategy to achieve stimuli-responsive and controlled drug release. We synthesized colloidal mesoporous silica (CMS) nanoparticles with ATTO633-labeled cysteine linked to the inner particle core via disulfide-bridges and characterized their cysteine release behavior after internalization into HuH7 cells by high-resolution fluorescence microscopy. Our study revealed that endosomal escape is a bottleneck for disulfide-linkage based drug release. Photochemical opening of the endosome leads to successful delivery of fluorescently labeled cysteine to the cytosol. KEYWORDS Colloidal mesoporous silica, nanoparticles, endosomal release, photosensitizer, redox-sensitive disulfide linker, drug delivery.

C

drugs in porous host materials. The concept of disulfidebound agents inside the pore system of mesoporous silica has been initially studied by Mortera et al.25 They showed that membrane-impermeable cysteine can be successfully delivered into cells. Cysteine delivery into cells promotes the synthesis of glutathione (GSH), which plays a central role in cell biology. A number of diseases including cancer as well as neurodegenerative and cardiovascular diseases are associated with GSH depletion and existing treatment strategies aim at promoting GSH synthesis by cysteine delivery.31 While Mortera et al.25 have demonstrated that cell-induced release can be based on disulfide bridges in mesoporous silica, the cellular uptake mechanism and the specific locus as well as the time point of reduction inside the cell remains unknown. For a further development and optimization of this promising approach, it is of particular interest to focus on the details of nanoparticle entry and subsequent events after entry into the cell. Such processes can be followed effectively with highly sensitive live-cell imaging techniques. These techniques can monitor the uptake of a single particle and its trafficking inside the cell.32-34 A well-known bottleneck for drug delivery into the cytosol of cells is the endosomal escape of macromolecular substances or drug conjugates.35 Because of the low reductive or even oxidative milieu in endosomes, contact to the cytosol is essential for release of disulfide-bound agents.36 The use of photoactive compounds to open the endosome is one possibility to overcome this challenge.37,38 Recently, some of us have used the disulphonated porphyrin derivative meso-tetraphenylporphine (TPPS2a) for the study of photoinduced endosomal release dynamics of DNA-containing polyplexes and dextranes in living cells.35 It was demonstrated that endosomal compartments can be damaged by activation of

ontrolled release of drugs from functionalized mesoporous materials has attracted great interest in recent years.1-6 Triggers used to affect the controlled release include enzymatic digestion of gatekeepers,7-10 temperature,7 competitive binding,11 light-irradiaton,12-19 changes in the pH value,20-23 or changes in the redox potential.24-28 An enzymatically triggered system was shown by us recently, using tightly bound avidin proteins to seal the pores.7 An opening of these protein caps was shown by the release of fluorescein upon proteolytic digestion of the attached avidin. Most of the systems use valves or caps at the pore openings of the porous silica surface.7-9,11,13,20,21,24 Furthermore, redox-sensitive, disulfide-based mechanisms were applied in the field of drug delivery.29 In the context of mesoporous silica, for example, poly(N-acryloxysuccinimide) was grafted on mesoporous silica nanoparticles and cross-linked by cystamine, resulting in a closure of the pores. Reopening was achieved by the addition of a reducing agent.30 As many therapeutic peptides carry thiol functions, and nucleic acid oligomers can be functionalized with such groups, redox-driven intracellular disulfide cleavage offers a high potential in drug delivery. This approach can also be transferred to the pore system of mesoporous hosts. Binding and releasing drug molecules from the inner volume of mesoporous silica nanoparticles allows one to avoid synthetically demanding molecular valve concepts while retaining the benefits of protecting sensitive or immunoresponsive

* To whom correspondence should be addressed. E-mail: (C.B.) Christoph.braeuchle@ cup.uni-muenchen.de; (T.B.) [email protected]. † These authors have equally contributed to the article. Received for review: 6/21/2010 Published on Web: 08/02/2010

© 2010 American Chemical Society

3684

DOI: 10.1021/nl102180s | Nano Lett. 2010, 10, 3684–3691

SCHEME 1. Synthesis of Core-Shell Functionalized Colloidal Mesoporous Silica

SCHEME 2. Attachment of ATTO633-Labeled Cysteine to the Core the Colloidal Mesoporous Silica

the membrane-bound photosensitizer to release the endosomal cargo. In the present work, photochemical release was used to study endosomal escape of colloidal mesoporous silica (CMS) nanoparticles in living HuH7 cells. Core-shell functionalized CMS nanoparticles were obtained using a recently developed strategy that allows integration of molecular functionalities at distinct locations within one mesoporous silica nanoparticle.39,40 We have synthesized CMS carrying a 3-mercaptopropyl-functionality at the inner pore walls of the core and 3-aminopropyl functionality on the outer particle surface. The outer surface was then labeled with N-hydroxysuccinimidyl-ATTO488 (ATTO488-NHS) fluorescent dye, while the inner core was functionalized with ATTO633-labeled cysteine by disulfide formation, leading to dual-color particles. The long-term intracellular integrity of the dual-color CMS nanoparticles measured by spinning disk confocal microscopy confirmed that endosomal escape is a bottleneck for drug delivery into cells. Endosomal release of the CMS nanoparticles was achieved by application of the photosensitizer TPPS2a and monitored by highly sensitive widefield fluorescence microscopy in living cells. The endosomal collapse was verified by the release of the fluid phase marker Alexa Fluor 488 Dextran (AFD). After endosomal collapse, release of nanoparticles, cleavage of the disulfide-bond and release of fluorescently labeled cysteine into the cytosol of living cells was observed. In summary, we show that real-time live-cell imaging of photochemical release is a powerful method to investigate in great detail the role and control of the endosomal escape. Our results demonstrate that endosomal escape is a limiting factor for the redox-triggered intracellular release of disulfide-bound cysteine from core-shell functionalized colloidal mesoporous silica (CMS). Results and Discussion. Synthesis and Bulk Characterization of Colloidal Mesoporous Silica. Core-shell functionalized CMS nanoparticles were synthesized according to our previously published sequential co-condensation method (Scheme 1).40 A mixture of TEOS and MPTES was hydrolyzed in an aqueous reaction mixture containing triethanolamine and cetyltrimethylammonium chloride. Thirty minutes after seed generation, a shell of pure silica was grown by adding four small amounts of TEOS (each 2.5 mol % of total Si content) in three minute steps. The mixture was stirred for another 30 min. Finally, a 1:1 mixture of TEOS: © 2010 American Chemical Society

TABLE 1. Overview of All Prepared Fluorescent CMS-Samples CMS sample name

core functionality

shell functionality

CysATTO633core-ATTO488shell N-ATTO633-labeled ATTO488 cysteine (disulfide-bridged) ATTO633coreATTO488shell CysATTO633core-NH2shell ATTO633core-NH2shell

ATTO633

ATTO488

N-ATTO633-labeled aminopropylcysteine (disulfide-bridged) ATTO633

aminopropyl-

APTES (2% of total silica) was added to the reaction. The resulting mixture was stirred for another 12 h at room temperature. After template extraction, colloidal mesoporous silica spheres featuring particle diameters of 80 nm (derived from transmission electron microscopy) were obtained (sample CMS-SHcore-NH2shell). To characterize the porous properties of the sample, nitrogen sorption measurements were performed, showing a Brunauer-Emmett-Teller surface area of 1160 m2 g-1 (isotherm can be found in the Supporting Information, Figure S-1). The pore size and volume was calculated to 3.8 nm and 0.93 cm3 g-1, respectively, according to nonlocal density functional theory. Subsequently, the amino-functionalized shell was labeled with ATTO488-NHS, yielding the sample CMS-SHcoreATTO488shell. After vigorous washing steps, the thiolfunctionalized core of the particles was activated with 2-2′-dithiopyridine (DTP) and further reacted with ATTO633labeled cysteine, resulting in a disulfide-bridged labeled cysteine attached to the inner core of the porous nanoparticle (sample CMS-CysATTO633core-ATTO488shell, Scheme 2). Additionally, a number of reference samples were prepared to perform all experiments described in the following. An overview of all prepared fluorescent CMS-samples is given in Table 1. A detailed particle synthesis protocol can be found in the Supporting Information (see Supporting Materials and Methods). The progress of the increasing functionalization was monitored by Raman spectroscopy (Figure 1). The relevant peaks in the spectrum are marked with asterisks. After the initial particle synthesis, the sample CMSSHcore-NH2shell (Figure 1a) shows the characteristic thiol vibra3685

DOI: 10.1021/nl102180s | Nano Lett. 2010, 10, 3684-–3691

FIGURE 2. Release of CysATTO633 from the sample CMSCysATTO633core-ATTO488shell. FIGURE 1. Raman spectra of the samples (a) CMS-SHcore-NH2shell, (b) DTP-activated CMS-SHcore-ATTO488shell, (c) CMS-CysATTO633coreATTO488shell, and (d) reduced CMS-SHcore-ATTO488shell.

Single-Particle Characterization of the Synthesized CMS Nanoparticles in Vitro. To follow the release of dyelabeled cysteine at a single-particle level by fluorescence microscopy, the functionalized CMS nanoparticles were sedimented on glass coverslips and incubated with 10 mM GSH solution. Movies of the nanoparticles were recorded before (t < 0 min) and after (t > 0 min) addition of GSH solution. The mean fluorescence intensity of the ATTO633 labeled particles and the background was extracted and the intensity was plotted versus time as shown in Figure 3. In Figure 3a, the relative CysATTO633 fluorescence intensity of CMS-CysATTO633core-ATTO488shell nanoparticles (black) and background (gray) is displayed. After GSH addition the CysATTO633 fluorescence of the background (Figure 3a, gray curve) increases up to a factor of 2.3 reaching a plateau after 8 min. The increase of CysATTO633 background fluorescence intensity indicates successful release of the dye from the functionalized CMS nanoparticles. Surprisingly, the CMS-CysATTO633core-ATTO488shell nanoparticles themselves (Figure 3a, black curves) also showed an increase in CysATTO633 fluorescence intensity. This increase by a factor of 2 or more occurred within two minutes and on a faster time scale than the increase in background fluorescence and was unexpected. Instead, we expected a decrease in the particle‘s CysATTO633 intensity after reductive dye release. However, the increase of nanoparticle-associated CysATTO633 intensity is easily explained by a dequenching effect of the pore-bound CysATTO633. The tight packing of the dye molecules inside the pores promotes a self-quenching of CysATTO633 similar to tightly packed octadecyl rhodamine B or calcein in liposomes.41 The release of dye lowers the dye concentration within the pores below the limit for selfquenching and the residual dye molecules start to fluoresce. The strong fluorescence of the nanoparticles also indicates that the disulfide-bound CysATTO633 is not completely released upon addition of GSH.

tion at 2584 cm-1. The next synthesis step was the activation of the thiol-functionalities using DTP (Figure 1b). The reaction is confirmed by Raman spectroscopy, by the emerging aromatic ring vibrations around 1000 cm-1 and the characteristic two bands for aromatic CN heterocycles at 1566 and 1580 cm-1. Additionally, the thiol vibration disappears upon disulfide formation. Further reaction of the activated thiols with dye-labeled cysteine leads to the removal of the aromatic vibrations in the Raman spectrum (Figure 1c), and there are still no free thiol functionalities visible in the spectrum. After the reduction of the newly formed disulfides with 10 mM GSH in a reference experiment, the Raman spectrum shows the signal of free thiols again (Figure 1d). The presence and absence of cysteine in the samples before and after the reduction are also confirmed with IR spectroscopy. The corresponding spectra can be found in the Supporting Information (Figure S-2). To examine the applicability of the system for GSH-responsive delivery, reference experiments were performed. For this purpose, the sample was transferred into our recently developed two-compartment fluorescence cuvette. The two compartments are separated by a dialysis membrane. While the CMS particles are too big to diffuse through the membrane, released, labeled cysteine can easily pass through the barrier and can be observed by fluorescence spectroscopy, applying an excitation wavelength of 633 nm. For the experiment, 1 mg of the sample in 200 µL of water was separated by the membrane from 3 mL of water in the other compartment. While no detectable amount of cysteine was released before the addition of GSH, the release starts immediately after creation of the reductive milieu (10 mM GSH in both compartments, Figure 2). One hour after the GSH addition, the concentration of cysteine remains stable. This maximum release was normalized to 100%. © 2010 American Chemical Society

3686

DOI: 10.1021/nl102180s | Nano Lett. 2010, 10, 3684-–3691

FIGURE 3. CysATTO633 release measured in vitro on a single-particle level. The mean fluorescence intensity of four single particles (black curves) on glass and background (gray curve) was extracted from a movie, normalized, and plotted versus time. The fluorescence intensity of ATTO633 was plotted for (a) redox-cleavable CMS-CysATTO633core-ATTO488shell and (b) noncleavable control CMS-ATTO633core-ATTO488shell nanoparticles. GSH was added at t ) 0 min. The movies were recorded at 642 nm illumination with an exposure time of 200 ms and a frame rate of 3.4 s-1.

The self-quenching effect of tightly packed ATTO633 in a constrained environment such as mesoporous silica has not been reported before. In the present study, this selfquenching effect of CysATTO633 permits a well-detectable readout for dye release with excellent signal-to-noise ratio. As a control, CMS-ATTO633core-ATTO488shell nanoparticles without cleavable disulfide linker were examined under similar conditions (Figure 3b). The fluorescence intensities of both the background and the particles remained constant and a dequenching effect was not observed after addition of 10 mM GSH (at t ) 0 min). This result confirms that only disulfide linker-bound dye is released by GSH and that the release is associated with a strong increase of nanoparticle fluorescence intensity. To summarize, we showed that release of disulfide bound CysATTO633 from CMS nanoparticles in 10 mM GSH can successfully be observed at a single-particle level. The release was accompanied by a dequenching of the ATTO633 fluorescence and this effect permits a sensitive readout for successful dye release. Live-Cell Imaging of HuH7 Cells Incubated with CMS Nanoparticles. To investigate whether the reductive milieu inside living cells is sufficient to induce dye release from CMS nanoparticles, we examined living cells exposed to CMSCysATTO633core-ATTO488shell nanoparticles for up to 2 days. After 24 and 48 h of exposure, confocal z-stacks of HuH7 cells were acquired by spinning disk confocal microscopy and the colocalization of CysATTO633core with ATTO488shell was evaluated. Successful cell entry and intracellular localization of the CMS nanoparticles was detected by their characteristic intracellular motion such as transport by motor proteins32-34 (see movies S-1A and S-1B in the Supporting Information) and the location within the z-stack. Transmission light images of the cells showed no morphological © 2010 American Chemical Society

signs of toxicity within our observation time. After 49 h, core-bound CysATTO633 and shell-bound ATTO488 signals of the intracellular particles were still colocalized. Additionally, fluorescence of free CysATTO633 in the cytoplasm was not detected. This indicates that within 49 h, the disulfidebound dye was not released. Two representative cells after 25 (Figure 4a-c) and 49 h (Figure 4d-f) of incubation are displayed. Fluorescence of the particle’s CysATTO633-core (red a + d) and ATTO488-shell (green b + e) was colocalized as indicated by the yellow signal in the merged image (yellow c + f), which was superimposed on the transmission light image of the cell. Our results indicate that CMS-CysATTO633core-ATTO488shell nanoparticles are taken up into HuH7 cells without detectable signs of toxicity. Reductive release of ATTO633-labeled cysteine from internalized particles was not detected within 49 h of incubation. Our data and previous studies show that CMS nanoparticles are taken up into and transported within endosomes.25 This implies that they are not accessible to the reductive milieu of the cytosol. Mortera et al. showed cytosolic fluorescence of reductively released cysteine.25 However, they did not show the release process itself. In our mechanistic study, we could not detect reductive release of ATTO633 labeled cysteine. Photochemically Induced Endosomal Release of CMS Nanoparticles in Living Cells. To overcome the endosomal membrane barrier that separates CMS nanoparticles from the cytosol and to follow the release at a single cell level in real time, we employed photoinduced endosomal release. In this method, the photosensitizer TPPS2a is incubated with the cells and incorporates into membranes via the endocytic pathway. TPPS2a is excited into its singlet state by 405 nm laser light, followed by intersystem crossing to its triplet 3687

DOI: 10.1021/nl102180s | Nano Lett. 2010, 10, 3684-–3691

FIGURE 4. Confocal microscopy of living HuH7 cells exposed to CMS-CysATTO633core-ATTO488shell nanoparticles. z-Projections are shown consisting of the overlay of three planes inside an HuH7 cell exposed to CMS-CysATTO633core-ATTO488shell nanoparticles for 25 h (a-c) and 49 h (d-f). The cell nucleus is indicated by a white circle and the outer white line represents the cell border. Fluorescence of the particles’ core (red a + d) and shell (green b + e) is highly colocalized as indicated by the yellow signal in the merged image (yellow c + f), which was superimposed on the transmitted light image of the cell. Scale bar: 10 µm.

state. This excited state is then quenched by triplet oxygen producing singlet oxygen. Singlet oxygen is able to oxidize unsaturated fatty acids, cholesterol, and amino acids and leads to a collapse of the endosomal membrane followed by release of the endosomal content into the cytosol. To label endosomes and monitor endosomal release, the fluid phase marker Alexa Fluor 488 Dextran (AFD) was added to the cells along with CMS-CysATTO633core-NH2shell nanoparticles and the photosensitizer. AFD is internalized by fluid-phase endocytosis.42 After 18-24 h of incubation, the cells were examined by wide-field fluorescence microscopy. All CMS nanoparticles that exhibit typical intracellular motion were found to be colocalized with AFD. This indicates successful internalization of the CMS nanoparticles into endosomes. The fluorescence intensity of the endosomes varied depend© 2010 American Chemical Society

ing on the amount of internalized AFD. However, the CysATTO633 fluorescence was quite weak, probably due to the self-quenching effect of the dye in the mesopores. Excitation of the photosensitizer was achieved by illumination of the sample with 405 nm laser for 1 min; this resulted in the termination of endosomal motion, as reported previously.35 Depending on the amount of photosensitizer incorporated in the endosomal membrane, the endosomes were ruptured within 1-4 min, leading to a spontaneous release of AFD into the cytoplasm as indicated by a sudden drop in endosomal AFD fluorescence and increase in cytosolic fluorescence. Concomitant with endosomal rupture and AFD release, the fluorescence intensity of the CMS-CysATTO633coreNH2shell nanoparticles increased due to the dequenching effect as described above and presented in Figure 5a. 3688

DOI: 10.1021/nl102180s | Nano Lett. 2010, 10, 3684-–3691

FIGURE 5. Photoinduced endosomal release of CMS-CysATTO633core-NH2shell and fluid phase marker AFD inside living HuH7 cells monitored by wide-field fluorescence microscopy. The cells were exposed to CMS-CysATTO633core-NH2shell overnight. (a) Intensity plot of three exemplary tracked endosomes (highlighted by small circles in b and c) over time. The fluorescence intensity of CysATTO633 (upper three curves) showed a sudden increase concomitant to the decrease in AFD fluorescence intensity (corresponding lower three curves) due to endosomal rupture and AFD-dye release. (b) Fluorescence microscopy image overlays of the CysATTO633 (red) and fluid phase marker AFD (green) signal at activation of the photosensitizer and (c) 4 min later. The cell nucleus is indicated with the large white circle. Scale bar: 10 µm.

FIGURE 6. Photoinduced endosomal release of CMS-ATTO633core-NH2shell and fluid phase marker AFD inside living HuH7 cells monitored by wide-field fluorescence microscopy. The cells were exposed to CMS-ATTO633core-NH2shell overnight. (a) Intensity plot of three exemplary tracked endosomes (highlighted by small circles in b and c) over time. The fluorescence intensity of ATTO633 (upper three curves) showed no or only slight fluorescence increase concomitant to the decrease in AFD fluorescence intensity (corresponding lower three curves) due to endosomal rupture and AFD-dye release. (b) Fluorescence microscopy image overlays of the ATTO633 (red) and fluid phase marker AFD (green) signal at activation of the photosensitizer and (c) 4 min later. The cell nucleus is indicated with the large white circle. Scale bar: 10 µm.

fluorescence intensity of CMS-CysATTO633core-NH2shell particles on glass (see Figure S-3 in the Supporting Information).

Directly after photosensitizer activation the endosomes show still predominantly AFD fluorescence (Figure 5b, depicted in green). Strikingly, only 4 min later the same endosomes show only CysATTO633 fluorescence (Figure 5c, depicted in red). Please note that the diffuse red fluorescence within the nucleus area (big, white circle) is due to out of focus particle fluorescence. With a size of 10 kDa, AFD can diffuse almost freely after endosomal release, and it is dispersed within the cytosol.43 In contrast, due to their large size and impaired motion in the crowded cell interior the nanoparticles remain at their location. The corresponding movie is available as movie S-2 in the Supporting Information. We also performed a control measurement showing that activated TPPS2a in solution has no influence on the © 2010 American Chemical Society

Asacontrolmeasurement,noncleavableCMS-ATTO633coreNH2shell nanoparticles without cysteine linker were incubated with AFD and the photosensitizer for 12-24 h. The internalized CMS-ATTO633core-NH2shell nanoparticles showed colocalization with AFD until endosomal rupture. At endosomal rupture, the relative fluorescence intensity of AFD showed a sudden drop whereas the ATTO633 fluorescence remained largely constant with a slight intensity increase by a factor of 1.25 (Figure 6a). This increase might be due to a small amount of unreacted dye inside the mesopores, which is released from the endosome after disruption of the endosomal membrane. The disruption of the endosomal mem3689

DOI: 10.1021/nl102180s | Nano Lett. 2010, 10, 3684-–3691

material is available free of charge via the Internet at http:// pubs.acs.org.

brane and release of AFD occurred within 4 min, and the CMS-ATTO633core-NH2shell nanoparticle fluorescence remained at the former endosomal regions (Figure 6c). Figure 6a displays the fluorescence intensity of AFD and CMSATTO633core-NH2shell nanoparticles of three representative endosomes (highlighted in Figure 6b,c with small circles). The corresponding movie S-3 is available in the Supporting Information. The missing increase in ATTO633 fluorescence intensity of noncleavable dye indicates that ATTO633 is maintained in the pores and did not escape into the cytosol as the disulfide-bound dye. Conclusions. Our results show that CMS-CysATTO633core nanoparticles were endocytosed by HuH7 cells without visible signs of toxicity. However, disulfide-bound dye in the pore system of internalized nanoparticles was not released within 49 h of observation. Inefficient endosomal escape of the nanoparticles is generally a bottleneck for molecular delivery into the cytoplasm. After photochemical rupture of the endosomes by means of a photosensitizer, CMSCysATTO633core nanoparticles successfully released disulfide-bound CysATTO633 into the cytoplasm showing that the reducing milieu of the cytoplasm is sufficient to cleave the cysteine linker. In case of noncleavable CMSATTO633core nanoparticles without cysteine linker, release of ATTO633 was not observed. In addition, we show for the first time that linkage of ATTO633 at high concentration in the pores of silica nanoparticles results in quenching of the ATTO633 fluorescence. Release of dye from the pores promotes a strong dequenching effect providing an intense fluorescence signal with excellent signal-to-noise ratio for single particle imaging.

REFERENCES AND NOTES (1) (2) (3) (4) (5) (6) (7) (8) (9)

(10) (11)

(12) (13) (14) (15) (16) (17) (18)

(19)

Acknowledgment. Support from the Nanosystems Initiative Munich (NIM), DFG (SFB 749), the Center for Integrated Protein Science Munich (CIPSM), and the Center for NanoScience (CeNS) is gratefully acknowledged. A.M.S. thanks the Elitenetzwerk Bayern. A.S. and A.M.S. thank the Ro¨mer Foundation for support. We also acknowledge the expert assistance of Monika Franke with the cell culture and Dr. Sergey Ivanchenko with the microscope setup.

(20) (21)

(22)

(23)

Supporting Information Available. Figure S-1: Nitrogen sorption isotherm of the sample CMS-SHcore-NH2shell. Figure S-2: Infrared spectroscopic characterization of the reductive release of cysteine from CMS material. Figure S-3: Influence of a TPPS2a solution on redox-cleavable CMS-CysATTO633coreNH2shell measured at a single particle level. Movie S-1A and S-1B: Confocal microscopy of living HuH7 cells exposed to CMS-CysATTO633core-ATTO488shell nanoparticles for 25 and 49 h. Movie S-2: Photoinduced endosomal release of CMSCysATTO633core-NH2shell and fluid phase marker Alexa Fluor 488 Dextran inside living HuH7 cells monitored by widefield fluorescence microscopy. Movie S-3: Photoinduced endosomal release of CMS-ATTO633core-NH2shell and fluid phase marker Alexa Fluor 488 Dextran inside living HuH7 cells monitored by wide-field fluorescence microscopy. This © 2010 American Chemical Society

(24) (25) (26) (27) (28)

(29) (30) (31)

3690

Vallet-Regi, M.; Balas, F.; Arcos, D. Angew. Chem., Int. Ed. 2007, 46, 7548–7558. Descalzo, A. B.; Martinez-Manez, R.; Sancenon, F.; Hoffmann, K.; Rurack, K. Angew. Chem., Int. Ed. 2006, 45, 5924–5948. Aznar, E.; Martinez-Manez, R.; Sancenon, F. Expert Opin. Drug Delivery 2009, 6, 643–655. Trewyn, B. G.; Slowing, I. I.; Giri, S.; Chen, H.-T.; Lin, V. S.-Y. Acc. Chem. Res. 2007, 40, 846–853. Slowing, I. I.; Vivero-Escoto, J. L.; Wu, C.-W.; Lin, V. S.-Y. Adv. Drug Delivery Rev. 2008, 60, 1278–1288. Saha, S.; Leung, K. C. F.; Nguyen, T. D.; Stoddart, J. F.; Zink, J. I. Adv. Funct. Mater. 2007, 17, 685–693. Schlossbauer, A.; Kecht, J.; Bein, T. Angew. Chem., Int. Ed. 2009, 48, 3092–3095. Patel, K.; Angelos, S.; Dichtel, W. R.; Coskun, A.; Yang, Y. W.; Zink, J. I.; Stoddart, J. F. J. Am. Chem. Soc. 2008, 130, 2382–2383. Bernardos, A.; Aznar, E.; Marcos, M. D.; Martinez-Manez, R.; Sancenon, F.; Soto, J.; Barat, J. M.; Amoros, P. Angew. Chem., Int. Ed. 2009, 48, 5884–5887. Park, C.; Kim, H.; Kim, S.; Kim, C. J. Am. Chem. Soc. 2009, 131, 16614–16615. Climent, E.; Bernardos, A.; Martinez-Manez, R.; Maquieira, A.; Marcos, M. D.; Pastor-Navarro, N.; Puchades, R.; Sancenon, F.; Soto, J.; Amoros, P. J. Am. Chem. Soc. 2009, 131, 14075–14080. Lu, J.; Choi, E.; Tamanoi, F.; Zink, J. I. Small 2008, 4, 421–426. Nguyen, T. D.; Leung, K. C. F.; Liong, M.; Liu, Y.; Stoddart, F.; Zink, J. I. Adv. Funct. Mater. 2007, 17, 2101–2110. Mal, N. K.; Fujiwara, M.; Tanaka, Y. Nature 2003, 421, 350–353. Mal, N. K.; Fujiwara, M.; Tanaka, Y.; Taguchi, T.; Matsukata, M. Chem. Mater. 2003, 15, 3385–3394. Zhu, Y.; Fujiwara, M. Angew. Chem., Int. Ed. 2007, 46, 2241–2244. Angelos, S.; Choi, E.; Voegtle, F.; De Cola, L.; Zink, J. I. J. Phys. Chem. C 2007, 111, 6589–6592. Liu, N. G.; Dunphy, D. R.; Atanassov, P.; Bunge, S. D.; Chen, Z.; Lopez, G. P.; Boyle, T. J.; Brinker, C. J. Nano Lett. 2004, 4, 551– 554. Aznar, E.; Casasus, R.; Garcia-Acosta, B.; Marcos, M. D.; MartinezManez, R.; Sancenon, F.; Soto, J.; Amoros, P. Adv. Mater. 2007, 19, 2228–2231. Nguyen, T. D.; Leung, K. C. F.; Liong, M.; Pentecost, C. D.; Stoddart, J. F.; Zink, J. I. Org. Lett. 2006, 8, 3363–3366. Angelos, S.; Khashab, N. M.; Yang, Y. W.; Trabolsi, A.; Khatib, H. A.; Stoddart, J. F.; Zink, J. I. J. Am. Chem. Soc. 2009, 131, 12912–12914. Casasus, R.; Marcos, M. D.; Martinez-Manez, R.; Ros-Lis, J. V.; Soto, J.; Villaescusa, L. A.; Amoros, P.; Beltran, D.; Guillem, C.; Latorre, J. J. Am. Chem. Soc. 2004, 126, 8612–8613. Casasus, R.; Climent, E.; Marcos, M. D.; Martinez-Manez, R.; Sancenon, F.; Soto, J.; Amoros, P.; Cano, J.; Ruiz, E. J. Am. Chem. Soc. 2008, 130, 1903–1917. Lai, C. Y.; Trewyn, B. G.; Jeftinija, D. M.; Jeftinija, K.; Xu, S.; Jeftinija, S.; Lin, V. S.-Y. J. Am. Chem. Soc. 2003, 125, 4451–4459. Mortera, R.; Vivero-Escoto, J.; Slowing, I. I.; Garrone, E.; Onida, B.; Lin, V. S.-Y. Chem. Commun. 2009, 22, 3219–3221. Hernandez, R.; Tseng, H. R.; Wong, J. W.; Stoddart, J. F.; Zink, J. I. J. Am. Chem. Soc. 2004, 126, 3370–3371. Nguyen, T. D.; Liu, Y.; Saha, S.; Leung, K. C. F.; Stoddart, J. F.; Zink, J. I. J. Am. Chem. Soc. 2007, 129, 626–634. Nguyen, T. D.; Tseng, H. R.; Celestre, P. C.; Flood, A. H.; Liu, Y.; Stoddart, J. F.; Zink, J. I. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 10029–10034. West, K. R.; Otto, S. Curr. Drug Discovery Technol. 2005, 2, 123– 160. Liu, R.; Zhao, X.; Wu, T.; Feng, P. J. Am. Chem. Soc. 2008, 130, 14418–14419. Pastore, A.; Federici, G.; Bertini, E.; Piemonte, F. Clin. Chim. Acta 2003, 333, 19–39. DOI: 10.1021/nl102180s | Nano Lett. 2010, 10, 3684-–3691

(32) Bausinger, R.; von Gersdorff, K.; Braeckmans, K.; Ogris, M.; Wagner, E.; Bra¨uchle, C.; Zumbusch, A. Angew. Chem., Int. Ed. 200645, 1568–1572. (33) de Bruin, K.; Ruthardt, N.; von Gersdorff, K.; Bausinger, R.; Wagner, E.; Ogris, M.; Bra¨uchle, C. Mol. Ther. 2007, 15, 1297– 1305. (34) Sauer, A. M.; de Bruin, K. G.; Ruthardt, N.; Mykhaylyk, O.; Plank, C.; Bra¨uchle, C. J. Controlled Release 2009, 137, 136–145. (35) de Bruin, K. G.; Fella, C.; Ogris, M.; Wagner, E.; Ruthardt, N.; Bra¨uchle, C. J. Controlled Release 2008, 130, 175–182. (36) Austin, C. D.; Wen, X.; Gazzard, L.; Nelson, C.; Scheller, R. H.; Scales, S. J. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 17987–17992. (37) Berg, K.; Selbo, P. K.; Prasmickaite, L.; Tjelle, T. E.; Sandvig, K.; Moan, J.; Gaudernack, G.; Fodstad, O.; Kjolsrud, S.; Anholt, H.;

© 2010 American Chemical Society

(38)

(39) (40) (41) (42) (43)

3691

Rodal, G. H.; Rodal, S. K.; Hogset, A. Cancer Res. 1999, 59, 1180– 1183. Berg, K.; Selbo, P. K.; Weyergang, A.; Dietze, A.; Prasmickaite, L.; Bonsted, A.; Engesaeter, B. O.; Angell-Petersen, E.; Warloe, T.; Frandsen, N.; Hogset, A. J. Microsc. (Oxford, U.K.) 2005, 218, 133–147. Kecht, J.; Schlossbauer, A.; Bein, T. Chem. Mater. 2008, 20, 7207– 7214. Cauda, V.; Schlossbauer, A.; Kecht, J.; Zu¨rner, A.; Bein, T. J. Am. Chem. Soc. 2009, 131, 11361–11370. Pollock, S.; Antrobus, R.; Newton, L.; Kampa, B.; Rossa, J.; Latham, S.; Nichita, N. B.; Dwek, R. A.; Zitzmann, N. FASEB J. 2010, 24, 1866–1878. Berlin, R. D.; Oliver, J. M. J. Cell. Biol. 1980, 85, 660–671. Dauty, E.; Verkman, A. S. J. Biol. Chem. 2005, 280, 7823–7828.

DOI: 10.1021/nl102180s | Nano Lett. 2010, 10, 3684-–3691